Inhomogeneity in Ethylammonium Nitrate–Acetonitrile Binary Mixtures

Jun 26, 2017 - The binary mixtures of the ionic liquid ethylammonium nitrate with acetonitrile have been studied by means of wide- and small-angle X-r...
0 downloads 10 Views 3MB Size
Subscriber access provided by EAST TENNESSEE STATE UNIV

Letter

Inhomogeneity in Ethylammonium Nitrate – Acetonitrile Binary Mixtures. Part I: The Highest “Low q Excess” So Far Reported Alessandro Mariani, Ruggero Caminiti, Fabio Ramondo, Giovanna Salvitti, Francesca Mocci, and Lorenzo Gontrani J. Phys. Chem. Lett., Just Accepted Manuscript • DOI: 10.1021/acs.jpclett.7b01244 • Publication Date (Web): 26 Jun 2017 Downloaded from http://pubs.acs.org on June 28, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry Letters is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 49

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

Inhomogeneity in Ethylammonium Nitrate – Acetonitrile Binary Mixtures. Part I: The Highest “Low q Excess” So Far Reported Alessandro Mariani*‡,[a] Ruggero Caminiti,[a][b] Fabio Ramondo,[c] Giovanna Salvitti,[c] Francesca Mocci,[d] Lorenzo Gontrani*[a] [a]

Dipartimento di Chimica, La Sapienza Università di Roma, Piazzale Aldo Moro 5, 00185 Rome – Italy [b]

Centro di Ricerca per le Nanotecnologie Applicate all’Ingegneria, Laboratorio per le

Nanotecnologie e le Nanoscienze, La Sapienza Università di Roma, Piazzale Aldo Moro 5, 00185 Rome – Italy [c]

Dipartimento di Scienze Fisiche e Chimiche, Università degli Studi dell’Aquila, Via Vetoio, L’Aquila I-67100 – Italy

[d]

Dipartimento di Scienze Chimiche e Geologiche, Università di Cagliari, S.S. 554 Km 4,500, I-09042 Monserrato - Italy

Corresponding Author *Alessandro Mariani [email protected] *Lorenzo Gontrani [email protected]

Alessandro Mariani’s new address is European Synchrotron Radiation Facility, 71 Avenue des

Martyrs, 38000 Grenoble

ACS Paragon Plus Environment

1

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 49

The binary mixtures of the ionic liquid ethylammonium nitrate with acetonitrile have been studied by means of Wide and Small Angle X-ray Scattering and via two different computational methods, namely classical molecular dynamics and DFT. Our results show how the local arrangement is directly linked to the long-range structure. Moreover, we found once again a similarity in the physicochemical behaviour of ethylammonium nitrate and water.

TOC GRAPHICS

Dissolving a co-solvent into an Ionic Liquid (IL) may have a variety of effects depending on the nature of the salt and on the added compound1–3. Taking Ethylammonium Nitrate (EAN) as a prototype, there is a wide range of literature concerning the effects that mixing has on its structure4–19. The molecular arrangement of neat EAN is shared with a number of other ILs and consists of a nano-segregation of two domains (polar and apolar) percolating each other8,20,21. This is the so-called sponge-like structure that is responsible for the Low q Peak (LqP) in the Small Angle X-ray Scattering (SAXS) patterns of these compounds22,23. Recently it has been reported that the addition of n-alcohols to some ILs leads to an unexpected feature in the extreme low q region of the SAXS pattern, which from now on will be termed “Low q Excess” (LqE)24,25. Greaves et al. observed that in some EAN-alcohol mixtures, some micellar-like structures could be found26. A confirmation of this behaviour came from Jiang et al. stating that when the alcohol

ACS Paragon Plus Environment

2

Page 3 of 49

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

alkyl chain is longer than twice that of the IL, then it is too big to be accommodated into the apolar domain thus forming a series of self-assembled structures where the cation acts as a cosurfactant25. In fact, they report the LqE for a variety of systems and always when the ratio between the chain lengths of the IL and the alcohol is less than 0.5. Nevertheless, this interpretation cannot explain why the same effect is found in EAN-methanol mixtures24, where one may expect a total mutual miscibility, since methanol is small enough to fit easily in EAN apolar domain. For that system, when LqE was first observed by Russina et al., an explanation in terms of ionic liquid molecules clustering was proposed to explain that odd feature, their EPSR (reverse Monte Carlo) simulations suggested that the feature below 0.3 Å-1 was generated by EAN clusters floating into a methanol sea. While the structure factor obtained by EPSR finely reproduced the experimental data, such a rigid interpretation is unlikely to be invoked for a liquid, macroscopically homogenous phase. More recent findings, on EAN+1-pentanol systems17, suggest that the origin of the LqE is a more complex critical phenomenon related to the density and concentration fluctuations experienced by the system at its incipient demixing. We have pointed out that the LqE is much more common than one may believe, and it is not limited to protic ionic liquid + n-alcohols systems. In our previous work18, the LqE was reported for EAN + 1,2-Dimethoxyethane and EAN + 1,4-Diaminobutane, both symmetric (nonamphiphilic) compounds. Our analysis suggested that the key quantity that may be directly linked to the LqE is the enthalpy of mixing. More precisely, an ionic liquid-molecular compound binary mixture will show the LqE when the enthalpy of mixing of the two components is larger than 30 kJ/mol for that composition. Based on our findings, the cause behind this unexpected feature in the SAXS patterns of some systems may be related to an incipient demixing of the sample driven by the lack of molecular affinity between the IL and the co-solvent. Thus LqE is

ACS Paragon Plus Environment

3

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 49

linked to incipient structural transition of the system. Classical Molecular Dynamics (MD) is the state-of -the-art technique to interpret, predict and understand complex systems, and it is widely used when ILs are considered3,27,28. Its limitation in neglecting electronic effects is overcome by the chance to simulate large systems, up to several thousands of atoms. A task that would be impossible for ab initio calculations. Nevertheless, good results have been achieved with MD in the field of ionic liquids, also when LqE is involved, like in our previous work. Anyway, DFT calculations of small clusters can give a much more reliable insight on the local short range molecular arrangement than the picture given from MD. For this reason, we have considered also some small ab initio models.

Figure 1. Schematic representation of (left) ethylammonium nitrate; (right) acetonitrile. In the present paper, we show how Acetonitrile (ACN) interacts with EAN when they are mixed together. ACN is structurally similar to methanol, but it is unable to donate hydrogen bonds, so it is the perfect candidate to check if hydrogen bond has a role in the LqE observation. Density fluctuations have been reported for aqueous mixtures of acetonitrile29–33, for which various groups have pointed out how water and acetonitrile do not homogeneously mix but they rather organize themselves into water-rich and acetonitrile-rich regions percolating each other. For this system, the LqE is also observed and it was analysed in terms of concentration fluctuation parameters, Kirkwood-Buff Integrals (KBIs)34 and/or Ornstein-Zernike (OZ) formalism. Each approach may take advantage of a range of computational techniques to get an atomistic insight into the structure. Nishikawa et al.29 used concentration fluctuation parameters and KBIs to interpreter their SAXS data. Basing their rationalization on Koga et al.35 work, they

ACS Paragon Plus Environment

4

Page 5 of 49

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

stated that the acetonitrile-water mixture is made of large clusters of the pure compounds, without trace of extended hydrogen bond network unlike the homologous water-methanol system. Subsequent studies involving several experimental and computational methods, confirmed this picture that is now firmly supported. In this work, we are presenting the molecular arrangement of EAN-ACN mixtures. We will interpret the experimental SAXS data using OZ formalism, then we will propose a computational model obtained by classical MD simulations. Finally, we will propose some small ab initio models to highlight the short-range molecular arrangement. Our observations are completely in line with those of Perron et al.36 who studied EAN-acetonitrile binary mixtures as a model for highly concentrated electrolyte solutions. The SAXS patterns collected as a function of composition, are plotted in Figure 2.

Figure 2. Small and Wide Angle X-ray scattering patterns of the systems studied. χEAN 0 (blue); 0.1 (red); 0.3 (green); 0.5 (purple); 0.7 (cyan); 0.9 (grey); 1 (black). It is evident how at low EAN concentrations, in the extreme low q region the scattered intensity has a very high and broad feature. Like all the previously reported systems, the LqE has its maximum intensity at the lower (analysed) EAN concentration17,18,24,25, and here we are presenting the highest LqE ever observed for this kind of system. Albeit MD models are not large enough (100 Å box side) to properly describe the range with q smaller than 0.15 Å-1, the

ACS Paragon Plus Environment

5

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 49

curves show a markedly steep rise approaching the lower q limit (see Supporting Information) and, as demonstrated elsewhere18, the initial slope of the LqE is reproduced by the models. The OZ formalism describes the diffracted intensity i(q), expressed in arbitrary units, as  =





+ 

(1)

where ξ is the correlation length of electron density fluctuations, bg is the background contribution and i(0) is the scattered intensity at q=0, in turn expressed in arbitrary units. To obtain the structure factor I(q) in absolute units, one must normalize i(q) using the formula  =  − ∑  

(2)

where xi and fi are the numerical concentration the scattering factor of atom i, respectively. I(0) is related to the isothermal compressibility κT by 0 =   

(3)

where, kB is the Boltzmann constant, T is the temperature in Kelvin and ρN is the number density of the system obtained by  =

 ∙

(4)

!"# ∙$%"# &!"# ∙$%'()*+,)-,.(

where NA is the Avogadro’s Number, ρ is the mass density of the system, χ represents the molar fraction and MW the molecular weight.

Table 1. Overview on some properties of the systems at 25 °C

ACS Paragon Plus Environment

6

Page 7 of 49

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

/01

i(0)[a] [A.U.]

[a] ξ

0.00

2.86e-6

0.10

 [b] [molecules*Å-3]

 [g*cm-3]

  [c] [TPa-1]

0.1002

0.0468

0.77590

1.15e3

1.03e-1

0.8978

0.0108

0.85719

9.23e8

0.31

4.24e-2

0.3496

0.0096

0.98347

1.88e8

0.50

1.38e-2

0.2035

0.0086

1.07155

5.17e7

0.71

1.56e-5

0.1002

0.0077

1.13984

5.57e4

0.90

1.14e-5

0.1001

0.0070

1.18966

4.08e4

1.00

9.58e-6

0.1001

0.0122

1.20920

1.91e4

[nm]

[a] from Equation 1; [b] from Equation 4; [c] from Equation 3

Applying equation 3 to compute the acetonitrile κT, a value of 1150 TPa-1 was found, in excellent agreement with 1154 TPa-1 reported by Grant-Taylor37, confirming the goodness of the procedure and the data. In our case the scattered intensity is much higher than the background, so the latter is negligible. Fitting the lowest q region of the experimental data with an OZ-like function, one is able to retrieve both i(0) and ξ. Another way to estimate i(0) comes from the Ornstein-Zernike plot. Plotting the reciprocal of the scattered intensity as a function of q2 a linear relationship is observed, which makes easier the extrapolation to q=0. The results of both the approaches are shown in Figure 3a, the fitting parameters for equation 1 are reported in Figure 3b together with κT obtained applying equation 3, and Table 1 summarizes the results.

ACS Paragon Plus Environment

7

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

a

Page 8 of 49

b

Figure 3. (a) Fitting of the SAXS pattern by an Ornstein-Zernike equation. Ornstein-Zernike plot in the inset. Same colours as in Figure 2. Experimental data (symbols); fitting (lines). (b) In the main frame, fitting parameters of the Ornstein-Zernike equation. i(0) (black squares, left Y axis); ξ (orange triangles, right Y axis). In the inset, isothermal compressibility of the systems. Lines are just a guide to the eye. Both fitting parameters for equation 1, clearly show two different trend regimes: i(0) steeply increases passing from neat acetonitrile to χEAN 0.1, and then linearly (in logarithmic scale) decreases as EAN is added to the mixture until a marked step between χEAN 0.5 and 0.7 breaks the linearity. Starting from χEAN 0.7 another linear correlation is found, but i(0) has decreased of several order of magnitude. The step separates the systems showing the LqE (χEAN ≤ 0.5) from the ones which do not do so (χEAN

≥ 0.7).

Although, the variation in ξ does not show a step, two

distinct linear regions are easily individuated. Once again these regions can be linked to the presence or absence of LqE. Here it appears clear that the objects responsible of the slope in the extreme low q region are rapidly formed as the IL is added to acetonitrile with a maximum dimension of ~0.9 nm at χEAN 0.1 and then exponentially decrease their size until χEAN 0.7 where the objects associated to the slope are molecular-sized and cannot (obviously) get smaller. A simple, yet powerful way to check the molecular interactions in a binary mixture is the study of

ACS Paragon Plus Environment

8

Page 9 of 49

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

the excess molar volume (VEX). This quantity is readily obtained from the experimental density measurements via 2 03 =

!4 $%4 ! $% (56

−7

!4 $%4 4

+

! $%

8

(5)

where MWx is the molecular weight of component 1 or 2, ρexp is the experimental density and ρx is the density of pure 1 or 2. The points obtained were then fitted using a Redlich-Kister function38, and the results are shown in Figure 4. The values of VEX for each temperature can be found in Table S1 in Supporting Information along with experimental densities. Fitting parameters are in Table S2 in Supporting Information.

Figure 4. Excess molar volume for ethylammonium nitrate-acetonitrile mixtures. 288 K (black circles); 293 K (white circles); 298 K (black reverse triangles); 303 K (white triangles); 308 K (black squares); 313 K (white squares); 318 K (black diamonds); 323 K (white diamonds); 328 K (black triangles). Data are fitted with a Redlich-Kister function (dashed line). VEX values are dependent from two different phenomena, namely molecular affinity and packing efficiency. In this case, the excess molar volume appears to be negative in the whole composition range for all the considered temperatures, meaning an overall favourable interaction between EAN and ACN, due to one or both the cited effects. Varying the temperature, the

ACS Paragon Plus Environment

9

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

position of the minimum remains almost constant at χ

EAN

Page 10 of 49

~0.35, suggesting that on average two

hydrogen bond donors on the cation are involved in interactions with as many acetonitrile molecules. Interestingly, increasing the temperature all the systems deviate more from ideality, thus enhancing the favourable interactions. It is known that coulombic forces and hydrogen bonds are weakened upon temperature rising, thus the stronger interaction must be due to increased packing efficiency. We have also computed the partial molar volumes differentiating the fitted excess molar volume values for each temperature. The results are reported in Figure 5, and in Table S1 in Supporting Information.

Figure 5. Partial molar volumes for ethylammonium nitrate-acetonitrile mixtures. 288 K (black); 293 K (red); 298 K (green); 303 K (orange); 308 K (blue); 313 K (purple); 318 K (cyan); 323 K (grey); 328 K (violet). EAN (solid line, right Y axis); ACN (dashed line, left Y axis). $ $ and 219 There are four distinct regions in both EAN and ACN partial molar volumes (201 $ respectively). The first one spans in the region 0